how likely is contagion in financial networks?...how likely is contagion in financial networks? ......

37
OFFICE OF FINANCIAL RESEARCH U.S. DEPARTMENT OF THE TREASURY Office of Financial Research Working Paper #0009 June 21, 2013 How Likely is Contagion in Financial Networks? Paul Glasserman, 1 and H. Peyton Young 2 1 Columbia Business School, Columbia University [email protected] 2 Department of Economics, University of Oxford, and Office of Financial Research, U.S. Department of Treasury, [email protected] The Office of Financial Research (OFR) Working Paper Series allows staff and their co-authors to disseminate preliminary research findings in a format intended to generate discussion and critical comments. Papers in the OFR Working Paper Series are works in progress and subject to revision. Views and opinions expressed are those of the authors and do not necessarily represent official OFR or Treasury positions or policy. Comments are welcome as are suggestions for improvements, and should be directed to the authors. OFR Working Papers may be quoted without additional permission. www.treasury.gov/ofr

Upload: others

Post on 21-Mar-2020

8 views

Category:

Documents


0 download

TRANSCRIPT

OFFICE OFFINANCIAL RESEARCHU.S. DEPARTMENT OF THE TREASURY

Office of Financial Research Working Paper #0009

June 21, 2013

How Likely is Contagion in Financial Networks?

Paul Glasserman,1 and H. Peyton Young2

1 Columbia Business School, Columbia University [email protected] 2 Department of Economics, University of Oxford, and Office of Financial Research, U.S. Department of Treasury, [email protected]

The Office of Financial Research (OFR) Working Paper Series allows staff and their co-authors to disseminate preliminary research findings in a format intended to generate discussion and critical comments. Papers in the OFR Working Paper Series are works in progress and subject to revision. Views and opinions expressed are those of the authors and do not necessarily represent official OFR or Treasury positions or policy. Comments are welcome as are suggestions for improvements, and should be directed to the authors. OFR Working Papers may be quoted without additional permission.

www.treasury.gov/ofr

How Likely is Contagion in Financial Networks?

Paul Glasserman∗ and H. Peyton Young†

Abstract

Interconnections among financial institutions create potential channels for contagion and amplifica-tion of shocks to the financial system. We estimate the extent to which interconnections increaseexpected losses, with minimal information about network topology, under a wide range of shockdistributions. Expected losses from network effects are small without substantial heterogeneity inbank sizes and a high degree of reliance on interbank funding. They are also small unless shocksare magnified by some mechanism beyond simple spillover effects; these include bankruptcy costs,fire sales, and mark-to-market revaluations of assets. We illustrate the results with data on theEuropean banking system.

Keywords: systemic risk, contagion, financial network

JEL: D85, G21

We thank Thomas Noe and Alireza Tahbaz-Salehi for constructive comments on an earlier draft.The views and opinions expressed are those of the authors and do not necessarily represent official OFRor Treasury positions or policy. Comments are welcome as are suggestions for improvements and shouldbe directed to the authors. OFR Working Papers may be quoted without additional permission.

∗Columbia Business School, Columbia University, [email protected]†Department of Economics, University of Oxford, and Office of Financial Research, U.S. Treasury, pey-

[email protected]

1 Introduction

The interconnectedness of the modern financial system is widely viewed as having been a key contributing

factor to the recent financial crisis. Due to the complex web of links between institutions, stresses to one

part of the system can spread to others, leading to a system-wide threat to financial stability. Specific

instances include the knock-on effects of the Lehman bankruptcy, potential losses to counterparties that

would have resulted from a failure of the insurance company AIG, and more recently the exposure of

European banks to the risk of sovereign default by some European countries. These and other examples

highlight concerns that interconnectedness could pose a significant threat to the stability of the financial

system.1 Moreover there is a growing body of research that shows how this can happen in a theoretical

sense.2

Although it is intuitively clear that interconnectedness has some effect on the transmission of shocks,

it is less clear whether and how it significantly increases the likelihood and magnitude of losses compared

to a financial system that is not interconnected. In this paper we propose a general framework for

analyzing this question. There are in fact many different types of networks connecting parts of the

financial system, including networks defined through ownership hierarchies, payment systems, derivatives

contracts, brokerage relationships, and correlations in stock prices, among other examples. We focus on

the network defined by liabilities between financial institutions. These payment obligations create the

most direct channel for the spread of losses.

It turns out that one can derive general bounds on the effects of this source of interconnectedness

with almost no information about the network topology: our bounds hold independently of the degree

distribution, amount of connectivity, node centrality, average path length, and so forth. The topology-

free property of our results is one of the main contributions of this work. Moreover, we impose probability

distributions on the shocks to the nodes and show that the same bounds hold for a wide range of

distributions, including beta, exponential, normal, and many others. This robustness is important

because detailed information about interbank liabilities is often unavailable and the exact form of the

shock distributions is subject to considerable uncertainty.

To model a network of payment obligations, we build on the elegant framework of Eisenberg and

Noe (2001). The model specifies a set of nodes that represent financial institutions together with the

obligations between them. Given an initial shock to the balance sheets of one or more nodes, one can

compute a set of payments that clear the network by solving a fixed-point problem. This framework is1See, for example, Bank of England 2011, International Monetary Fund 2012, and Office of Financial Research 2012.2See in particular Allen and Gale (2000), Upper and Worms (2002), Degryse and Nguyen (2004), Goodhart, Sunirand,

and Tsomocos (2004), Elsinger, Lehar, and Summer (2006), Allen and Babus (2009), Gai and Kapadia (2010), Gai,Haldane, and Kapadia (2011), Haldane and May (2011), Upper (2011), Allen and Carletti (2011), Georg (2011), Rogersand Veraart (2012), Acemoglu, Ozdaglar, and Tahbaz-Salehi (2013), and Elliott, Golub, and Jackson (2013).

1

very useful for analyzing how losses propagate through the financial system. A concrete example would

be delinquencies in mortgage payments: if some fraction of a banks mortgages are delinquent and it has

insufficient reserves to cover the shortfall, then it will be unable to pay its creditors in full, who may

be unable to pay their creditors in full, and so forth. The original shortfall in payments can cascade

through the system, causing more and more banks to default through a domino effect. The Eisenberg-Noe

framework shows how to compute a set of payments that clear the network, and it identifies which nodes

default as a result of an initial shock to the system. The number and magnitude of such defaults depend

on the topological structure of the network, and there is now a substantial literature characterizing those

structures that tend to propagate default or alternatively that tend to dampen it (Gai and Kapadia 2010,

Gai, Haldane, and Kapadia 2011, Haldane and May, 2011, Acemoglu, Ozdaglar, and Tahbaz-Salehi 2013,

and Elliott, Golub, and Jackson 2013).

Much of this literature proceeds by examining the effects of fixed shocks applied to particular nodes

rather than fully specifying the distribution that generates the shocks. In this paper we analyze the

probability of contagion and the expected losses generated by contagion when the joint distribution of

shocks is given. We then apply the framework to answer the following questions about the impact of

network effects. First, how likely is it that a given set of banks will default due to contagion from another

node, as compared to the likelihood that they default from direct shocks to their own assets? Second,

how much does the network increase the probability and magnitude of losses compared to a situation

where there are no connections?

To compare systems with and without interconnections, we proceed as follows. First, we define our

nodes to be financial institutions that borrow and lend on a significant scale, which together with their

obligations to one another constitute the financial network. In addition, such institutions borrow and

lend to the nonfinancial sector, which is composed of investors, households, and nonfinancial firms. We

compare this system to one without connections that is constructed as follows. We remove all of the

obligations between the financial nodes while keeping their links with the nonfinancial sector unchanged.

We also keep node equity values as before by creating, for each node, a fictitious outside asset (or liability)

whose value equals the net value of the connections at that node that were removed. We then apply

the same shock distributions to both systems, with the shocks to real assets originating in the external

sector and the fictitious assets (if any) assumed to be impervious to shocks. We can ascertain how much

the network connections contribute to increased defaults and losses by comparing the outcomes in the

two systems.

One might suppose that the comparison hinges on what shock distribution we use, but this turns

out not to be the case: we show how to compute general bounds on the increased losses attributable to

network contagion that hold under a wide variety of distributions. The bounds also hold whether the

2

shocks are independent or positively associated and thus capture the possibility that institutions have

portfolios that are exposed to common shocks (see for example Caccioli et al. 2012).

Two key findings emerge from this analysis. First we compute the probability that default at a given

node causes defaults at other nodes (via network spillovers), and compare this with the probability that

all of these nodes default by direct shocks to their outside assets with no network transmission. We

derive a general formula that shows when the latter probability is larger than the former, in which case

we say that contagion is weak. A particular implication is that contagion is always weak unless there is

substantial heterogeneity in node sizes as measured by their claims outside the financial sector. More

generally, contagion will tend to be weak unless the originating node is large, highly leveraged, and –

crucially – has a relatively high proportion of its obligations to other financial institutions as opposed to

the nonfinancial sector. Second, the analysis shows that the total additional losses generated by network

spillover effects are surprisingly small under a wide range of shock distributions for plausible values of

model parameters. Both of these results are consistent with the empirical and simulation literature on

network stress testing, which finds that contagion is quite difficult to generate through the interbank

spillover of losses (Degryse and Nguyen 2004, Elsinger, Lehar, and Summer 2006, Furfine 2003, Georg

2011, Nier et al. 2007). Put differently, our results show that contagion through spillover effects becomes

most significant under the conditions described in Yellen (2013), when financial institutions inflate their

balance sheets by increasing leverage and expanding interbank claims backed by a fixed set of real assets.

These results do not imply that all forms of network contagion are unimportant; rather they show that

simple spillover or “domino” effects have only a limited impact at realistic levels of payment obligations

between banks. This leads us to examine other potential sources of contagion. The prior literature

has focused on the role of fire sales, that is, the dumping of assets on the market in order to cover

losses.3 Here we shall focus on two alternative mechanisms that are more immediate extensions of the

simple spillover mechanism in the Eisenberg-Noe model and thus allow a more immediate comparison:

bankruptcy costs and losses of confidence.

Bankruptcy costs magnify the costs associated with default both directly, through costs like legal

fees, and indirectly through delays in payments to creditors and disruptions to the provision of financial

intermediation services necessary to the real economy. We model these effects in reduced form through a

multiplier on losses when a node defaults. This approach allows us to estimate how much the probability

of contagion, and the expected losses induced by contagion, increase as a function of bankruptcy costs. A

somewhat surprising finding is that bankruptcy costs must be quite large in order to have an appreciable

impact on expected losses as they propagate through the network.

A second mechanism that we believe to be of greater importance is crises of confidence in the credit3See Shleifer and Vishny (2011) for a survey and Cifuentes, Ferrucci, and Shin (2005) for an extension of the Eisenberg-

Noe framework with fire sales.

3

quality of particular firms. If a firm’s perceived ability to pay declines for whatever reason, then so does

the market value of its liabilities. In a mark-to-market regime this reduction in value can spread to other

firms that hold these liabilities among their assets. In other words, the mere possibility (rather than the

actuality) of a default can lead to a general and widespread decline in valuations, which may in turn

trigger actual defaults through mark-to-market losses.4 This is an important phenomenon in practice:

indeed it has been estimated that mark-to-market losses from credit quality deterioration exceeded losses

from outright defaults in 2007-2009.5

We capture this idea by re-interpreting the Eisenberg-Noe framework as a valuation model rather

than as a clearing model. Declines in confidence about the ability to pay at some nodes can spread

to other nodes through a downward revaluation of their assets. This mechanism shows how a localized

crisis of confidence can lead to widespread losses of value. Our analysis suggests that this channel of

contagion is likely to be considerably more important than simple domino or spillover effects.

The rest of the paper is organized as follows. In Section 2 we present the basic Eisenberg-Noe

framework and illustrate its operation through a series of simple examples. In Section 3 we introduce

shock distributions explicitly. We then compare the probability that a given set of nodes default from

simultaneous direct shocks to their outside assets, with the probability that they default indirectly by

contagion from some other node. In Section 4 we examine the expected loss in value that is attributable

to network contagion using the comparative framework described above. We show that one can obtain

useful bounds on the losses attributable to the network with almost no knowledge of the specific network

topology and under very general assumptions about the shock distributions. In Section 5 we introduce

bankruptcy costs and show how to extend the preceding analysis to this case. Section 6 examines the

effects of a deterioration in confidence at one or more institutions, such as occurred in the 2008-09

financial crisis. We show how such a loss of confidence can spread through the entire system due to

mark-to-market declines in asset values. In the Appendix we illustrate the application of these ideas to

the European banking system using data from European Banking Authority (2011).

2 Measuring systemic risk

2.1 The Eisenberg-Noe framework

The network model proposed by Eisenberg and Noe (2001) has three basic ingredients: a set of n nodes

N = {1, 2, ..., n}, an n × n liabilities matrix P = (pij) where pij ≥ 0 represents the payment due from

4This mechanism differs from a bank run, which could also be triggered by a loss of confidence. Mark-to-market lossesspread when a lender continues to extend credit, whereas a run requires withdrawal of credit. In the seminal frameworkof Diamond and Dybvig (1983), a run is triggered by a demand for liquidity rather than a concern about credit quality.

5According to the Basel Committee on Banking Supervision, for example, roughly two thirds of losses attributed tocounterparty credit risk were due to mark-to-market losses and only about one third of losses were due to actual defaults.See http://www.bis.org/press/p110601.htm.

4

wi

bi

ci wj

wk

ijp

kip

Figure 1 Figure 1: Node i has an obligation pij to node j, a claim pkj on node k, outside assets ci, and outsideliabilities bi, for a net worth of wi.

node i to node j, pii = 0, and a vector c = (c1, c2, ..., cn) ∈ Rn+ where ci ≥ 0 represents the value

of outside assets held by node i in addition to its claims on other nodes in the network. Typically ci

consists of cash, securities, mortgages and other claims on entities outside the network. In addition each

node i may have liabilities to entities outside the network; we let bi ≥ 0 denote the sum of all such

liabilities of i, which we assume have equal priority with i’s liabilities to other nodes in the network.

The asset side of node i’s balance sheet is given by ci +∑

j 6=i pji, and the liability side is given by

pi = bi +∑

j 6=i pij . Its net worth is the difference

wi = ci +∑j 6=i

pji − pi. (1)

The notation associated with a generic node i is illustrated in Figure 1. Inside the network (indicated

by the dotted line), node i has an obligation pij to node j and a claim pki on node k. The figure also

shows node i’s outside assets ci and outside liabilities bi. The difference between total assets and total

liabilities is the node’s net worth wi.

Observe that i’s net worth is unrestricted in sign; if it is nonnegative then it corresponds to the

book value of i’s equity. We call this “book value” because it is based on the nominal or face value of

the liabilities pji, rather than on “market” values that reflect the nodes’ ability to pay. These market

values depend on other nodes’ ability to pay conditional on the realized value of their outside assets.

To be specific, let each node’s outside assets be subjected to a random shock that reduces the value of

its outside assets, and hence its net worth. These are shocks to “fundamentals” that propagate through

the network of financial obligations. Let Xi ∈ [0, ci] be a random shock that reduces the value of i’s

outside assets from ci to ci−Xi. After the shock, i’s net worth has become wi−Xi. Let F (x1, x2, ..., xn)

be the joint cumulative distribution function of these shocks; we shall consider specific classes of shock

distributions in the next section. (We use Xi to denote a random variable and xi to denote a particular

5

5

10

50

55

10

50

55

10

5

100 150

10

5 5

55

50

55

50

10

Figure 2

5

10

50

55

10

50

55

10

5

100 150

10

5 5

55

50

55

50

10 y

y

y

y

Figure 3

(a) (b)

Figure 2: Two network examples.

realization.)

To illustrate the effect of a shock, we consider the numerical example in Figure 2(a), which follows

the notational conventions of Figure 1. In particular, the central node has a net worth of 10 because it

has 150 in outside assets, 100 in outside liabilities, and 40 in liabilities to other nodes inside the network.

A shock of magnitude 10 to the outside assets erases the central node’s net worth, but leaves it with just

enough assets (140) to fully cover its liabilities. A shock of magnitude 80 leaves the central node with

assets of 70, half the value of its liabilities. Under a pro rata allocation, each liabilitity is cut in half,

so each peripheral node receives a payment of 5, which is just enough to balance each peripheral node’s

assets and liabilities. Thus, in this case, the central node defaults but the peripheral nodes do not. A

shock to the central node’s outside assets greater than 80 would reduce the value of every node’s assets

below the value of its liabilities.

Figure 2(b) provides a more complex version of this example in which a cycle of obligations of size

y runs through the peripheral nodes. To handle such cases, we need the notion of a clearing vector

introduced by Eisenberg and Noe (2001).

The relative liabilities matrix A = (aij) is the n× n matrix with entries

aij ={

pij/pi, if pi > 00, if pi = 0 (2)

Thus aij is the proportion of i’s obligations owed to node j. Since i may also owe entities in the external

sector,∑

j 6=i aij ≤ 1 for each i, that is, A is row substochastic.6

Given a shock realization x = (x1, x2, ..., xn) ≥ 0, a clearing vector p(x) ∈ Rn+ is a solution to the

system

pi(x) = pi ∧ (∑

j

pj(x)aji + ci − xi)+. (3)

6The row sums are all equal to 1 in Eisenberg and Noe (2001) because bi = 0 in their formulation.

6

As we shall subsequently show, the clearing vector is unique if the following condition holds: from every

node i there exists a chain of positive obligations to some node k that has positive obligations to the

external sector. (This amounts to saying that A has spectral radius less than 1.) We shall assume that

this condition holds throughout the remainder of the paper.

2.2 Mark-to-market values

The usual way of interpreting p(x) is that it corresponds to the payments that balance the realized assets

and liabilities at each node given that: i) debts take precedence over equity; and ii) all debts at a given

node are written down pro rata when the net assets at that node (given the payments from others), is

insufficient to meet its obligations. In the latter case the node is in default, and the default set is

D(p(x)) = {i : pi(x) < pi(x)}. (4)

However, a second (and, in our setting, preferable) way of interpreting p(x) is to see (3) as a mark-to-

market valuation of all assets following a shock to the system. The nominal value pi of node i’s liabilities

is marked down to pi(x) as a consequence of the shock x, including its impact on other nodes. As in our

discussion above of Figure 2(a), after marking-to-market, node i’s net worth is reduced from (1) to

ci − xi +∑j 6=i

pji(x)− pi(x). (5)

The reduction in net worth reflects both the direct effect of the shock component xi and the indirect

effects of the full shock vector x. Note, however, that this is a statement about values; it does not require

that the payments actually be made at the end of the period. Under this interpretation p(x) provides a

consistent re-valuation of the assets and liabilities of all the nodes when a shock x occurs.

As shown by Eisenberg and Noe, a solution to (3) can be constructed iteratively as follows. Given a

realized shock vector x define the mapping Φ : Rn+ → Rn

+ as follows:

∀i, Φi(p) = pi ∧ (∑

j

pjaji + ci − xi)+. (6)

Starting with p0 = p let

p1 = Φ(p0), p2 = Φ(p1), ... (7)

This iteration yields a monotone decreasing sequence p0 ≥ p1 ≥ p2... . Since it is bounded below it has

a limit p′, and since Φ is continuous p′ satisfies (3). Hence it is a clearing vector.

We claim that p′ is in fact the only solution to (3). Suppose by way of contradiction that there is

another clearing vector, say p′′ 6= p′. As shown by Eisenberg and Noe, the equity values of all nodes

must be the same under the two vectors, that is,

p′A + (c− x)− p′ = p′′A + (c− x)− p′′.

7

Rearranging terms it follows that

(p′′ − p′)A = p′′ − p′, where p′′ − p′ 6= 0.

This means that the matrix A has eigenvalue 1, which is impossible because under our assumptions A

has spectral radius less than 1.

3 Estimating the probability of contagion

Systemic risk can be usefully decomposed into two components: i) the probability that a given set of

nodes D will default and ii) the loss in value conditional on D being the default set. This decomposition

allows us to distinguish between two distinct phenomena: contagion and amplification. Contagion occurs

when defaults by some nodes trigger defaults by other nodes through a domino effect. Amplification

occurs when contagion stops but the losses among defaulting nodes keep escalating because of their

indebtedness to one another. Roughly speaking the first effect corresponds to a “widening” of the crisis

whereas the second corresponds to a “deepening” of the crisis. In this section we shall examine the

probability of contagion; the next section deals with the amplification of losses due to network effects.

To estimate the probability of contagion we shall obviously need to make assumptions about the

distribution of shocks. We claim, however, that we can estimate the relative probability of contagion

versus simultaneous default with virtually no information about the network structure and relatively

weak conditions on the shock distribution.

To formulate our results we shall need the following notation. Let βi = pi/(bi + pi) be the proportion

of i’s liabilities to other entities in the financial system.7 We can assume that βi > 0 , since otherwise

node i would effectively be outside the financial system. Recall that wi is i’s initial net worth (before a

shock hits), and ci is the initial value of its outside assets. The shock Xi to node i takes values in [0, ci].

We shall assume that wi > 0, since otherwise i would already be insolvent. We shall also assume that

wi ≤ ci, since otherwise i could never default directly through losses in its own outside assets. Define

the ratio λi = ci/wi ≥ 1 to be the leverage of i’s outside assets. (This is not the same as i’s overall

leverage, which in our terminology is the ratio of i’s total assets to i’s net worth.)

3.1 A general bound on the probability of contagion

Proposition 1. Suppose that only node i receives a shock, so that Xj = 0 for all j 6= i. Suppose that

no nodes are in default before the shock. Fix a set of nodes D not containing i. The probability that the7Elliott, Golub, and Jackson (2013) have a similar measure which they call the level of integration. More broadly, Shin

(2012) discusses the reliance of banks on wholesale funding as a contributor to financial crises, and βi measures the degreeof this reliance in our setting.

8

shock causes all nodes in D to default is at most

P (Xi ≥ wi + (1/βi)∑j∈D

wj). (8)

Moreover, contagion from i to D is impossible if∑j∈D

wj/wi > βi(λi − 1). (9)

The condition in (9) states that contagion from i to D is impossible if the total net worth of the

nodes in D is sufficiently large relative to the net worth of i weighted by the exposure of the financial

system to node i, as measured by βi, and the vulnerability of i as measured by its leverage. A similar

interpretation applies to (8).

Before proceeding to the proof, we illustrate the impossibility condition in (9) through the network

in Figure 2(a). The central node is node i, meaning that the shock affects its outside assets, and the

remaining nodes comprise D. The relevant parameter values are βi = 2/7, λi = 15, and the net worths

are as indicated in the figure. The left side of (9) evaluates to 2 and the right side to 4, so the condition

is violated, and, indeed, we saw earlier that contagion is possible with a shock greater than 80. However,

a modification of the network that raises the sum of the net worths of the peripheral nodes above 40

makes contagion impossible. This holds, for example, if the outside liabilities of every peripheral node

are reduced by more than 5, or if the outside liabilities of a single peripheral node are reduced by more

than 20. This example also illustrates that (9) is tight in the sense that if the reverse strict inequality

holds, then contagion is possible in this example with a sufficiently large shock.

Proof of Proposition 1. Let D(x) ≡ D be the default set resulting from the shock vector X, whose

coordinates are all zero except for Xi. By assumption i causes other nodes to default, hence i itself must

default, that is, i ∈ D. To prove (8) it suffices to show that

βi(Xi − wi) ≥∑

j∈D−{i}

wj ≥∑j∈D

wj . (10)

The second inequality in (10) follows from the assumption that no nodes are in default before the shock

and the fact that we must have D ⊆ D − {i} for all nodes in D to default.

For the first inequality in (10), define the shortfall at node j to be the difference sj = pj − pj . From

(3) we see that the vector of shortfalls s satisfies

s = (sA− w + X)+ ∧ p.

By (4) we have sj > 0 for j ∈ D and sj = 0 otherwise. We use a subscript D as in sD or AD to restrict

a vector or matrix to the entries corresponding to nodes in the set D. Then the vector of shortfalls at

9

the nodes of D satisfies

sD ≤ sDAD − wD + XD, (11)

hence

XD − wD ≥ sD(ID −AD). (12)

The vector sD is strictly positive in every coordinate. From the definition of βj we also know that the

jth row sum of ID −AD is at least 1− βj . Hence,

sD(ID −AD) · 1D ≥∑j∈D

sj(1− βj) ≥ si(1− βi). (13)

From (11) it follows that the shortfall at node i is at least as large as the initial amount by which i

defaults, that is,

si ≥ Xi − wi > 0. (14)

From (12)–(14) we conclude that∑j∈D

(Xj − wj) = Xi − wi −∑

j∈D−{i}

wj ≥ si(1− βi) ≥ (Xi − wi)(1− βi). (15)

This establishes (10) and the first statement of the proposition. The second statement follows from

the first by recalling that the shock to the outside assets cannot exceed their value, that is, Xi ≤ ci.

Therefore by (8) the probability of contagion is zero if ci < wi + (1/βi)∑

j∈D wj . Dividing through by

wi we see that this is equivalent to the condition∑

j∈D wj/wi > βi(λi − 1). �

The preceding proposition relates the probability of contagion from a given node i to the net worth

of the defaulting nodes in D relative to i’s net worth. The bounds are completely general and do not

depend on the distribution of shocks or on the topology of the network. The critical parameters are βi,

the degree to which the triggering node is indebted to the financial sector, and λi, the degree of leverage

of i’s outside assets.

The Appendix gives estimates of these parameters for large European banks, based on data from

stress tests conducted by the European Banking Authority (2011). Among the 50 largest of these banks

the average of the λi is 24.9, the average of our estimated βi is 14.9%, and the average of the products

βi(λi − 1) is 3.2. Proposition 1 implies that contagion from a “typical” bank i cannot topple a set of

banks D if the net worth of the latter is more than 3.2 times the net worth of the former, unless there

are additional channels of contagion.8

8On average, commercial banks in the United States are leveraged only about half as much as European banks, andtheir values of βi are somewhat smaller (Federal Reserve Release H.8, Assets and Liabilities of Commercial Banks in theUnited States, 2012). This suggests that contagion is even less likely in the US financial sector than in Europe.

10

3.2 Contagion with proportional shocks

We can say a good deal more if we impose some structure on the distribution of shocks. The notion

that contagion from i to D is weak, described informally in Section 1, can now be made precise by the

condition

P (Xi ≥ wi + (1/βi)∑j∈D

wj) ≤ P (Xi > wi)∏j∈D

P (Xj > wj). (16)

The expression on the left bounds the probability that these nodes default solely through contagion from

i, while the expression on the right is the probability that the nodes in D default through independent

direct shocks. Contagion is weak if the latter probability is at least as large as the former. The assumption

of independent direct shocks is somewhat unrealistic: in practice one would expect the shocks to different

nodes to be positively associated. In this case, however, the probability of default from direct shocks is

even larger, hence weak contagion covers this situation as well.

Let us assume that the losses at a given node i scale with the size of the portfolio ci. Let us also

assume that the distribution of these relative losses is the same for all nodes, and independent among

nodes. Then there exists a distribution function H : [0, 1] → [0, 1] such that

F (x1, ..., xn) =∏

1≤i≤n

H(xi/ci). (17)

Beta distributions provide a flexible family with which to model the distribution of shocks as a fraction

of outside assets. We work with beta densities of form

hp,q(y) =yp−1(1− y)q−1

B(p, q), 0 ≤ y ≤ 1, p, q ≥ 1, (18)

where B(p, q) is a normalizing constant. The subset with p = 1 and q > 1 has a decreasing density and

seems the most realistic, but (18) is general enough to allow a mode anywhere in the unit interval. The

case q = 1, p > 1 has an increasing density and could be considered “heavy-tailed” in the sense that it

assigns greater probability to greater losses, with losses capped at 100 percent of outside assets.9

Theorem 1. Assume the shocks are i.i.d. beta distributed as in (18) and that the net worth of every

node is initially nonnegative. Let D be a nonempty subset of nodes and let i /∈ D. Contagion from i to

D is impossible if ∑j∈D

wj > wiβi(λi − 1) (19)

and it is weak if ∑j∈D

wj ≥ wiβi

∑j∈D

(λi − 1)/λj . (20)

9Bank capital requirements under Basel II and III standards rely on a family of loss distributions derived from aGaussian copula model. As noted by Tasche (2008) and others, these distributions can be closely approximated by betadistributions.

11

As noted after Proposition 1, the condition in (19) states that contagion from i to D is impossible if

the total net worth of the nodes in D is sufficiently large relative to the net worth of i weighted by the

exposure of the financial system to node i and the vulnerability of i as measured by its leverage. The

condition in (20) compares the total net worth of D relative to that of i with the leverage of i relative

to that of the nodes in D. With other parameters held constant, increasing the relative net worth of

D makes contagion weaker in the sense that it strengthens the inequality; increasing the leverage of i

relative to that of nodes in D has the opposite effect. Importantly, the two effects are mediated by βi,

which measures the exposure of the financial system to node i — a lower βi makes D less vulnerable

to i and makes D less sensitive to the degree of leverage at i. Thus, (20) captures the effects of equity

levels, leverage ratios, and the degree of reliance on interbank lending on the risk of contagion.

By recalling that λj = cj/wj we can rewrite (20) in the equivalent form∑j∈D

cjλ−1j /

∑j∈D

λ−1j ≥ ciβi(1− λ−1

i ). (21)

Written this way, the condition states that contagion from i to D is weak if the average size of the nodes

in D weighted by their inverse leverage ratios (their capital ratios) is sufficiently large relative to i; on

the right side of the inequality, ciβi measures the financial system’s exposure to node i’s outside assets,

and the factor (1− λ−1i ) is greater when node i is more highly leveraged. Inequality (21) is thus harder

to satisfy, and D more vulnerable to contagion from i, if the large nodes in D are more highly leveraged,

if node i draws more of its funding from the financial system, or if node i is more highly leveraged.

Through (21), a key implication of Theorem 1 is that without some heterogeneity, contagion will be

weak irrespective of the structure of the interbank network:

Corollary 1. Assume that all nodes hold the same amount of outside assets ck ≡ c. Under the assump-

tions of Theorem 1, contagion is weak from any node to any other set of nodes.

Proof. This follows from the fact that βi(1 − λ−1i ) < 1; hence, if ck ≡ c, inequality (21) holds for all i

and D. �

In the Appendix we apply our framework to the 50 largest banks in the stress test data from the

European Banking Authority. It turns out that contagion is weak in a wide variety of scenarios. In

particular, we analyze the scenario in which one of the five largest European banks (as measured by

assets) topples two other banks in the top 50. We find that the probability of such an event is less than

the probability of direct default unless the two toppled banks are near the bottom of the list of 50.

In the example of Figure 2(a), with node i the central node, the left side of (20) evaluates to 20, and

the right side evaluates to 16 because βi = 2/7, λi = 15, and each of the peripheral nodes has λj = 10.

Thus, contagion is weak.

12

Proof of Theorem 1. Proposition 1 implies that contagion is weak from i to D if

P (Xi ≥ wi + (1/βi)∑j∈D

wj) ≤ P (Xi > wi)∏j∈D

P (Xj > wj). (22)

On the one hand this certainly holds if wi + (1/βi)∑

j∈D wj > ci, for then contagion is impossible. In

this case we obtain, as in (9), ∑j∈D

wj/wi > βi(λi − 1). (23)

Suppose on the other hand that (wi + (1/βi)∑

j∈D wj) ≤ ci. By assumption the relative shocks Xk/ck

are independent and beta distributed as in (18). In the uniform case p = q = 1, (22) is equivalent to

[1− (wi/ci + (1/βici)∑j∈D

wj)] ≤ (1− wi/ci)∏j∈D

(1− wj/cj), (24)

We claim that (24) implies (22) for the full family of beta distributions in (19). To see why, first observe

that the cumulative distribution Hp,q of hp,q satisfies

1−Hp,q(y) = Hq,p(1− y).

Hence (22) holds if

Hq,p(1− wi/ci − (1/βici)∑j∈D

wj) ≤ Hq,p(1− wi/ci)∏j∈D

Hq,p(1− wj/cj). (25)

But (25) follows from (24) because beta distributions with p, q ≥ 1 have the submultiplicative property

Hq,p(xy) ≤ Hq,p(x)Hq,p(y), x, y ∈ [0, 1].

(See Proposition 4.1.2 of Wirch 1999; the application there has q ≤ 1, but the proof remains valid for

q ≥ 1. The inequality can also be derived from Corollary 1 of Ramos Romero and Sordo Diaz 2001.) It

therefore suffices to establish (25), which is equivalent to

(1/βici)∑j∈D

wj ≥ (1− wi/ci)(1−∏j∈D

(1− wj/cj)). (26)

Given any real numbers θj ∈ [0, 1] we have the inequality∏j

(1− θj) ≥ 1−∑

j

θj . (27)

Hence a sufficient condition for (26) to hold is that

(1/βici)∑j∈D

wj ≥ (1− wi/ci)(∑j∈D

wj/cj). (28)

13

After rearranging terms and using the fact that λk = ck/wk for all k, we obtain (20). This concludes

the proof of Theorem 1. �

From the argument following (25), it is evident that the same result holds if the shocks to each node

j are distributed with parameters pj,qj in (18) with pi ≤ minj∈D pj and qi ≥ maxj∈D qj .

As a further illustration of Theorem 1, suppose the nodes in D are numbered 2, . . . ,m and suppose

node i = 1 receives a shock. Further suppose the outside assets are ordered c1 ≥ c2 ≥ · · · ≥ cm; because

shocks are proportional to outside assets, the assumption that the node with the largest ck receives the

shock maximizes the chances of contagion to the other nodes.

Corollary 2. If c1 ≥ c2 ≥ · · · ≥ cm, then contagion from node 1 to nodes 2, . . . ,m is weak if c2 ≥

β1(c1−w1) and cj ≥ (cj−1−wj−1), j = 2, . . . ,m. Contagion is impossible if, in addition, c2−cm+wm >

β1(c1 − w1).

This is a direct consequence of (24) and (26), hence we omit the details and comment on the in-

terpretation. The lower bounds on the cj ensure that the potential spillovers from other nodes cannot

push the full set of nodes into default regardless of the network topology. Viewing the conditions in the

corollary as lower bounds on the wj suggests minimum capital requirements to ensure the resilience of

the system, based on the relative sizes of banks.

Of course these results do not say that the network structure has no effect on the probability of

contagion; indeed there is a considerable literature showing that it does (see among others Haldane

and May 2011, Gai and Kapadia 2011, Georg, 2011). Rather it shows that in quite a few situations

the probability of contagion will be lower than the probability of direct default, absent some channel of

contagion beyond spillovers through payment obligations. We have already mentioned bankruptcy costs,

fire sales, and mark-to-market losses as amplifying mechanisms. The models of Demange (2012) and

Acemoglu, Ozdaglar, and Tahbaz-Salehi (2013) generate greater contagion by making debts to financial

institutions subordinate to other payment obligations. With priority given to outside payments, shocks

produce greater losses within the network. In practice, bank debt and bank deposits are owned by both

financial institutions and non-financial firms and individuals, so characterizing seniority based on the

type of lender is problematic.

3.3 Contagion with truncated shocks

In this section we shall show that the preceding results are not an artifact of the beta distribution:

similar bounds hold for a variety of shock distributions. Under the beta distribution the probability is

zero that a node loses all of its outside assets. One could easily imagine, however, that the probability

of this event is positive. This situation can be modeled as follows. Let Xoi ≥ 0 be a primary shock

14

(potentially unbounded in size) and let Xi = ci ∧ Xoi be the resulting loss to i’s outside assets. For

example Xoi might represent a loss of income from an employment shock that completely wipes out i’s

outside assets. Assume that the primary shocks have a joint distribution function of form

F o(xo1, ..., x

on) =

∏1≤i≤n

Ho(xoi /ci), (29)

where Ho is a distribution function on the nonnegative real line. In other words we assume that the

shocks are i.i.d. and that a given shock xoi affects every dollar of outside assets ci equally. (We shall

consider a case of dependent shocks after the next result.)

In general a random variable with distribution function G and density g is said to have an increasing

failure rate (IFR) distribution if g(x)/(1 − G(x)) is an increasing function of x. Examples of IFR

distributions include all normal, exponential, and uniform distributions and, more generally, all log-

concave distributions. Observe that truncating the shock can put mass at ci and thus assign positive

probability to a total loss of outside assets.

Theorem 2. Assume the primary shocks are i.i.d. and IFR-distributed, and that the net worth of every

node is initially nonnegative. Let D be a nonempty subset of nodes and let i /∈ D. Contagion from i to

D is impossible if ∑j∈D

wj > wiβi(λi − 1)

and it is weak if ∑j∈D

wj > wiβi

∑j∈D

λi/λj . (30)

Corollary 3. Assume that all nodes hold the same amount of outside assets ci ≡ c. Under the assump-

tions of Theorem 2, contagion is weak from any node to any other set of nodes.

This is immediate upon rewriting (30) as∑j∈D

cjλ−1j /

∑j∈D

λ−1j ≥ βici.

Proof of Theorem 2 . Through relabeling, we can assume that the source node for contagion is i = 1

and that the infected nodes are D = {2, 3, . . . ,m}. By Proposition 1 we know that contagion is weak

from 1 to D if

P (X1 > w1 + (1/β1)∑

2≤j≤m

wj) ≤ P (X1 > w1)P (X2 > w2) · · ·P (Xm > wm). (31)

15

Since X1 = c1 ∧ Xo1 , the left-hand side is zero when w1 + (1/β1)

∑2≤j≤m wj > c1. Thus contagion is

impossible if ∑2≤j≤m

wj/w1 > β1(λ1 − 1). (32)

Let us therefore assume that w1 + (1/β1)∑

2≤j≤m wj ≤ c1. Define the random variables Yi = Xoi /ci.

Weak contagion from 1 to D holds if

P (Y1 > w1/c1 + (1/β1c1)∑

2≤j≤m

wj) ≤ P (Y1 > w1/c1)P (Y2 > w2/c2) · · ·P (Ym > wm/cm)

= P (Y1 > w1/c1)P (Y1 > w2/c2) · · ·P (Y1 > wm/cm). (33)

where the latter follows from the assumption that the Yi are i.i.d. By assumption Y1 is IFR, hence

P (Y1 > s + t|Y1 > s) ≤ P (Y1 > t) for all s, t ≥ 0. (See for example Barlow and Proschan 1975, p.159.)

It follows that

P (Y1 >∑

1≤k≤m

wk/ck)

= P (Y1 > w1/c1)P (Y1 > w1/c1 + w2/c2|Y1 > w1/c1)

· · ·P (Y1 > w1/c1 + w2/c2 + · · ·+ wm/cm|Y1 > w1/c1 + w2/c2 + · · ·+ wm−1/cm−1)

≤ P (Y1 > w1/c1)P (Y1 > w2/c2) · · ·P (Y1 > wm/cm)

Together with (33) this shows that contagion from 1 to D is weak provided that

P (Y1 ≥ w1/c1 + (1/β1c1)∑

2≤j≤m

wj) ≤ P (Y1 >∑

1≤k≤m

wk/ck). (34)

This clearly holds if

w1/c1 + (1/β1c1)∑

2≤j≤m

wj ≥∑

1≤k≤m

wk/ck, (35)

which is equivalent to

(1/β1c1)∑

2≤j≤m

wj ≥∑

2≤j≤m

wj/cj =∑

2≤j≤m

λ−1j . (36)

Since c1 = λ1w1, we can re-write (36) as∑2≤j≤m

wj/w1 ≥ β1λ1

∑2≤j≤m

λ−1j . (37)

We have therefore shown that if contagion from 1 to D = {2, 3, . . . ,m} is possible at all, then (37) is a

sufficient condition for weak contagion. �

From (37), we see that a simple sufficient condition for weak contagion is cj ≥ β1c1, j = 2, . . . ,m,

and the condition∑m

j=2 wj > β1(c1 − w1) makes contagion impossible.

16

The assumption of independent shocks to different nodes is conservative. If we assume that direct

shocks to different nodes are positively dependent, as one would expect in practice, the bounds in

Theorems 1 and 2 will be lower and the relative likelihood of default through contagion even smaller.

A particularly simple case arises when the primary shocks Xi are independent with a negative expo-

nential distribution of form

f(xi) = µe−µxi , xi ≥ 0. (38)

If we further assume that c1 = · · · = cm ≡ c and β1 = 1, then the two probabilities compared in (33) are

equal, both evaluating to exp(−µ∑

j∈D wj/c). In this sense, the exponential distribution is a borderline

case in which the probability of a set of defaults from a single shock is roughly equal to the probability

from multiple independent shocks. We say “roughly” because the left side of (33) is an upper bound on

the probability of default through contagion, and in practice the βi are substantially smaller than 1. In

the example of Figure 2(a), we have seen that contagion from the central node requires a shock greater

than 80, which has probability exp(−80µ) under an exponential distribution. For direct defaults, it

suffices to have shocks greater than 5 at the peripheral nodes and a shock greater than 10 at the central

node, which has probability exp(−30µ) given i.i.d. exponential shocks.

If the primary shocks have a Pareto-like tail, meaning that

P (Xi > x) ∼ ax−µ (39)

for some positive constants a and µ (or, more generally, a regularly varying tail), then the probability

that a single shock will exceed∑

j∈D wj/c will be greater than the probability that the nodes in D

default through multiple independent shocks, at least at large levels of the wj . However, introducing

some dependence can offset this effect, as we now illustrate. To focus on the issue at hand, we take c = 1

and βi = 1.

To consider a specific and relatively simple case, let Y1, . . . , Ym be independent random variables,

each distributed as tν , the Student t distribution with ν > 2 degrees of freedom. Let Y1, . . . , Ym have

a standard multivariate Student t distribution with tν marginals.10 The Yj are uncorrelated but not

independent. To make the shocks positive, set Xj = Y 2j and Xj = Y 2

j . Each Xj and Xj has a Pareto-like

tail that decays with a power of ν/2.

Proposition 2. With independent shocks Xj,

P (Xi >m∑

j=1

wj) ≥m∏

j=1

P (Xj > wj)

10More explicitly, (Y1, ..., Ym) has the distribution of (Z1, ..., Zm)/p

χ2ν/ν, where the Zi are independent standard normal

random variables and χ2v has a chi-square distribution with v degrees of freedom and is independent of the Zi.

17

for all sufficiently large wj, j = 1, . . . ,m. With dependent shocks

P (Xi >m∑

j=1

wj) ≤ P (Xj > wj , j = 1, . . . ,m)

for all wj ≥ 0, j = 1, . . . ,m.

Proof of Proposition 2 . The first statement follows from applying (39) to both sides of the inequality.

The second statement is an application of Bound II for the F distribution on p.1196 of Marshall and

Olkin (1974). �

Thus, even with heavy-tailed shocks, we may find that default of a set of nodes through contagion

from a single shock is less likely than default through direct shocks to individual nodes if the shocks are

dependent.

4 Amplification of losses due to network effects

The preceding analysis dealt with the impact of default by a single node (the source) on another set of

nodes (the target). Here we shall examine the impact of shocks on the entire system, including multiple

and simultaneous defaults. To carry out such an analysis, we need to have a measure of the total systemic

impact of a shock. There appears to be no commonly accepted measure of systemic impact in the prior

literature. Eisenberg and Noe (2001) suggest that it is the number of waves of default that a given shock

induces in the network. Other authors have suggested that the systemic impact should be measured

by the aggregate loss of bank capital; see for example Cont, Moussa, and Santos (2010). Still others

have proposed the total loss in value of only those nodes external to the financial sector, i.e. firms and

households.

Here we shall take the systemic impact of a shock to be the total loss in value summed over all nodes,

including nodes corresponding to financial entities as well as those representing firms and households.

This measure is easily stated in terms of the model variables. Given a shock realization x, the total

reduction in asset values is

∑i xi + S(x) where S(x) =

∑i(pi − pi(x)). (40)

The term |x| =∑

i xi is the direct loss in value from reductions in payments by the external sector.

The term S(x) is the indirect loss in value from reductions in payments by the nodes to other nodes and

to the external sector. An overall measure of the riskiness of the system is the expected loss in value

L =∫

(|x|+ S(x))dF (x). (41)

The question we wish to examine is what proportion of these losses can be attributed to connections

between institutions as opposed to characteristics of individual banks. To analyze this issue let x be

18

a shock and let D = D(x) be the set of nodes that defaults given x. Under our assumptions this set

is unique because the clearing vector is unique. To avoid notational clutter we shall suppress x in the

ensuing discussion.

As in the proof of Propostion 1, define the shortfall in payments at node i to be si = pi − pi, where

p is the clearing vector. By definition of D,

si > 0 for all i ∈ Dsi = 0 for all i /∈ D

(42)

Also as in the proof of Proposition 1, let AD be the |D| × |D| matrix obtained by restricting the relative

liabilities matrix A to D, and let ID be the |D| × |D| identity matrix. Similarly let sD be the vector of

shortfalls si corresponding to the nodes in D, let wD be the corresponding net worth vector defined in

(1), and let xD be the corresponding vector of shocks. The clearing condition (3) implies the following

shortfall equation, provided no node is entirely wiped out — that is, provided si < pi, for all i:

sDAD − (wD − xD) = sD. (43)

Allowing the possibility that some si = pi, the left side is an upper bound on the right side. Recall that

AD is substochastic, that is, every row sum is at most unity. Moreover, by assumption, there exists a

chain of obligations from any given node k to a node having strictly positive obligations to the external

sector. It follows that limk→∞AkD = 0D, hence ID −AD is invertible and

[ID −AD]−1 = ID + AD + A2D + .... (44)

From (43) and (44) we conclude that

sD = (xD − wD)[ID + AD + A2D + ...]. (45)

Given a shock x with resulting default set D = D(x), define the vector u(x) ∈ Rn+ such that

uD(x) = [ID + AD + A2D + ...] · 1D , ui(x) = 0 for all i 6∈ D. (46)

Combining (40), (45), and (46) shows that total losses given a shock x can be written in the form

L(x) =∑

i

(xi ∧ wi) +∑

i

(xi − wi)ui(x). (47)

The first term represents the direct losses to equity at each node and the second term represents the

total shortfall in payments summed over all of the nodes. The right side becomes an upper bound on

L(x) if si = pi for some i ∈ D(x).

We call the coefficient ui = ui(x) the depth of node i in D = D(x). The rationale for this terminology

is as follows. Consider a Markov chain on D with transition matrix AD. For each i ∈ D, ui is the expected

19

number of periods before exiting D, starting from node i.11 Expression (46) shows that the node depths

measure the amplification of losses due to interconnections among nodes in the default set.

We remark that the concept of node depth is dual to the notion of eigenvector centrality in the

networks literature (see for example Newman 2010). To see the connection let us restart the Markov

chain uniformly in D whenever it exits D. This modified chain has an ergodic distribution proportional

to 1D · [ID + AD + A2D + ...], and its ergodic distribution measures the centrality of the nodes in D. It

follows that node depth with respect to AD corresponds to centrality with respect to the transpose of

AD.

Although they are related algebraically, the two concepts are quite different. To see why let us return

to the example of Figure 2(b). Suppose that node 1 (the central node) suffers a shock x1 > 80. This

causes all nodes to default, that is, the default set is D′ = {1, 2, 3, 4, 5}. Consider any node j > 1. In

the Markov chain described above the expected waiting time to exit the set D′, starting from node j, is

given by the recursion uj = 1 + βjuj , which implies

uj = 1/(1− βj) = 1 + y/55. (48)

From node 1 the expected waiting time satisfies the recursion

u1 = 1(100/140) + (1 + uj)(40/140). (49)

Hence

u1 = 9/7 + 2y/385. (50)

Comparing (48) and (50) we find that node 1 is deeper than the other nodes (u1 > uj) for 0 ≤ y < 22

and shallower than the other nodes for y > 22. In contrast, node 1 has lower eigenvector centrality than

the other nodes for all y ≥ 0 because it cannot be reached directly from any other node.

The magnitude of the node depths in a default set can be bounded as follows. In the social networks

literature a set D is said to be α-cohesive if every node in D has at least α of its obligations to other

nodes in D, that is,∑

j∈D aij ≥ α for every i ∈ D (Morris, 2000). The cohesiveness of D is the maximum

such α, which we shall denote by αD. From (46) it follows that

∀i ∈ D, ui ≥ 1/(1− αD). (51)

Thus the more cohesive the default set, the greater the depth of the nodes in the default set and the

greater the amplification of the associated shock.

Similarly we can bound the node depths from above. Recall that βi is the proportion of i’s obligations

to other nodes in the financial system. Letting βD = max{βi : i ∈ D} we obtain the upper bound,11Liu and Staum (2012) show that the node depths can be used to characterize the gradient of the clearing vector p(x)

with respect to the asset values.

20

assuming βD < 1,

∀i ∈ D, ui ≤ 1/(1− βD) . (52)

The bounds in (51) and (52) depend on the default set D, which depends on the shock x. A uniform

upper bound is given by

∀i, ui ≤ 1/(1− β+) where β+ = maxβi, (53)

assuming β+ < 1.

We are now in a position to compare the expected systemic losses in a given network of interconnec-

tions, and the expected systemic losses without such interconnections. As before, fix a set of n nodes

N = {1, 2, ..., n}, a vector of outside assets c = (c1, c2, ..., cn) ∈ Rn+ and a vector of outside liabilities

b = (b1, b2, ..., bn) ∈ Rn+. Assume the net worth wi of node i is nonnegative before a shock is realized.

Interconnections are determined by the n× n liabilities matrix P = (pij).

Let us compare this situation with the following: eliminate all connections between nodes, that is,

let P o be the n × n matrix of zeroes. Each node i still has outside assets ci and outside liabilities bi.

To keep their net worths unchanged, we introduce “fictitious” outside assets and liabilities to balance

the books. In particular if ci − bi < wi we give i a new class of outside assets in the amount c′i =

wi− (ci− bi). If ci− bi > wi we give i a new class of outside liabilities in the amount b′i = wi− (ci− bi).

We shall assume that these new assets are completely safe (they are not subject to shocks), and that

the new liabilities have the same priority as all other liabilities.

Let F (x1, ..., xn) be a joint shock distribution that is homogeneous in assets, that is, F (x1, ..., xn) =

G(x1/c1, ..., xn/cn) where G is a symmetric c.d.f. (Unlike in the preceding results on contagion we do

not assume that the shocks are independent.) We say that F is IFR if its marginal distributions are

IFR; this is equivalent to saying that the marginals of G are IFR. Given F , let L be the expected total

losses in the original network and let Lo be the expected total losses when the connections are removed

as described above.

Theorem 3. Let N(b, c, w, P ) be a financial system and let No be the analogous system with all the

connections removed. Assume that the shock distribution is homogeneous in assets and IFR. Let β+ =

maxi βi < 1 and let δi = P (Xi ≥ wi). The ratio of expected losses in the original network to the expected

losses in No is at mostL

Lo≤ 1 +

∑δici

(1− β+)∑

ci. (54)

Proof. By assumption the marginals of F are IFR distributed. A general property of IFR distributions

is that “new is better than used in expectation,” that is,

∀i ∀wi ≥ 0, E[Xi − wi|Xi ≥ wi] ≤ E[Xi] (55)

21

(Barlow and Proschan 1975, p.159). It follows that

∀i ∀wi ≥ 0, E[(Xi − wi)+] ≤ P (Xi ≥ wi)E[Xi] = δiE[Xi]. (56)

By (47) we know that the total expected losses L can be bounded as

L ≤∑

i

E[Xi ∧ wi] + E[∑

i

(Xi − wi)ui(X)]. (57)

From (53) we know that ui ≤ 1/(1 − β+) for all i; furthermore we clearly have Xi − wi ≤ (Xi − wi)+

for all i. Therefore

L ≤∑

i

E[Xi ∧ wi] + (1− β+)−1∑

i

E[(Xi − wi)+]. (58)

From this and (56) it follows that

L ≤∑

i

E[Xi ∧ wi] + (1− β+)−1∑

δiE[Xi]

≤∑

i

E[Xi] + (1− β+)−1∑

δiE[Xi]. (59)

When the network connections are excised, the expected loss is simply the expected sum of the shocks,

that is, Lo=∑

i E[Xi]. By the assumption of homogeneity in assets we know that E[Xi] ∝ ci for all i.

We conclude from this and (59) that

L/Lo ≤ 1 +∑

δici

(1− β+)∑

ci.

This completes the proof of Theorem 3. �

Theorem 3 shows that the increase in losses due to network interconnections will be very small unless

β+ is close to 1 or the default rates of some banks (weighted by their outside asset base) is large.

Moreover this statement holds even when the shocks are dependent, say due to common exposures, and

it holds independently of the network structure.

Some might argue that write-downs of purely financial obligations should not be counted as part of

the systemic loss. This is tantamount to truncating node depth at 1 and thus leads to an even smaller

upper bound in Theorem 3.

To illustrate the result concretely, consider the EBA data. In this case β+ = max{βi} = 0.43 where

the maximum occurs for Dexia bank. Regulatory restrictions on assets imply that any given bank is

unlikely to fail due to direct default; indeed current regulatory standards seek to make the direct default

probability of any individual bank smaller than 0.1% per year. Let us make the more conservative

assumption that the probability of default is at most 1%, that is, max{δi} < 0.01. If this standard is

met, then no matter what the interconnections might be among the banks in the data set, we conclude

22

from Theorem 3 that the additional expected loss attributable to the network is at most .01/(1− .43),

which is about 1.7%. In other words the actual network of connections cannot increase the expected

losses by very much under the assumptions of the theorem.

5 Bankruptcy costs

In this section and the next we enrich the basic framework by incorporating additional mechanisms

through which losses propagate from one node to another. We begin by adding bankruptcy costs. The

equilibrium condition (3) implicitly assumes that if node i’s assets fall short of its liabilities by 1 unit,

then the total claims on node i are simply marked down by 1 unit below the face value of pi. In practice,

the insolvency of node i is likely to produce deadweight losses that have a knock-on effect on the shortfall

at node i and at other nodes.

Cifuentes, Ferrucci, and Shin (2005), Battiston et al. (2012), Cont, Moussa, and Santos (2010),

and Rogers and Veraart (2012) all use some form of liquidation cost or recovery rate at default in their

analyses. Cifuentes, Ferrucci, and Shin (2005) distinguish between liquid and illiquid assets and introduce

an external demand function to determine a recovery rate on illiquid assets. Elliott, Golub, and Jackson

(2013) attach a fixed cost to bankruptcy. Elsinger, Lehar, and Summer (2006) present simulation results

illustrating the effect of bankruptcy costs but without an explicit model. The mechanism we use appears

to be the simplest and the closest to the original Eisenberg-Noe setting, which facilitates the analysis of

its impact on contagion.

5.1 Shortfalls with bankruptcy costs

In the absence of bankruptcy costs, when a node fails its remaining assets are simply divided among

its creditors. To capture costs of bankruptcy that go beyond the immediate reduction in payments, we

introduce a multiplier γ ≥ 0 and suppose that upon a node’s failure its assets are further reduced by

γ

pi − (ci +∑j 6=i

pjaji − xi)

, (60)

up to a maximum reduction at which the assets are entirely wiped out. This approach is analytically

tractable and captures the fact that large shortfalls are considerably more costly than small shortfalls,

where the firm nearly escapes bankruptcy. The term in square brackets is the difference between node

i’s obligations pi and its remaining assets. This difference measures the severity of the failure, and the

factor γ multiplies the severity to generate the knock-on effect of bankruptcy above and beyond the

immediate cost to node i’s creditors. We can think of the expression in (60) as an amount of value

destroyed or paid out to a fictitious bankruptcy node upon the failure of node i.

23

The resulting condition for a payment vector replaces (3) with

pi = pi ∧

(1 + γ)(ci +∑j 6=i

pjaji − xi)− γpi

+

. (61)

Written in terms of shortfalls si = pi − pi, this becomes

si = (1 + γ)[∑j 6=i

sjaji − wi + xi]+ ∧ pi. (62)

Here we see explicitly how the bankruptcy cost factor γ magnifies the shortfalls.

Let Φsγ denote the mapping from the vector s on the right side of (62) to the vector s on the left

side, and let Φpγ similarly denote the mapping of p defined by (61). We have the following:

Proposition 3. (a) For any γ ≥ 0, the mapping Φsγ is monotone increasing, bounded, and continuous

on Rn+, and the mapping Φp

γ is monotone decreasing, bounded, and continuous. It follows that Φsγ has a

least fixed point s and Φpγ has a greatest fixed point p = p − s. (b) For any s, Φs

γ(s) is increasing in γ,

and for any p, Φpγ(p) is decreasing in γ. Consequently, the set of default nodes under the minimal s and

maximal p is increasing in γ. The fixed point s (and p) is unique if (1 + γ)A has spectral radius less

than 1.

Proof of Proposition 3. Part (a) follows from the argument in Theorem 1 of Eisenberg and Noe (2001).

For part (b), write vi = ci + (pA)i − xi and observe that

Φpγ(p)i =

{vi − γ(pi − vi), vi < pi;pi otherwise,

from which the monotonicity in γ follows. The maximal fixed-point is the limit of iterations of Φpγ

starting from p by the argument in Section 3 of Eisenberg and Noe (2001). If γ1 ≤ γ2, then the iterates

of Φpγ1

are greater than those of Φpγ2

, so their maximal fixed-points are ordered the same way. But then

the set of nodes i for which pi < pi at the maximal fixed-point for γ1 must be contained within that for

γ2. The same argument works for Φsγ . Uniqueness follows as in the case without bankruptcy costs. �

This result confirms that bankruptcy costs expand the set of defaults (i.e., increase contagion) result-

ing from a given shock realization x while otherwise leaving the basic structure of the model unchanged.

In what follows, we examine the amplifying effect of bankruptcy costs conditional on a default set D,

and we then compare costs with and without network effects. The factor of 1 + γ already points to the

amplifying effect of bankruptcy costs, but we can take the analysis further.12

Suppose, for simplicity, that the maximum shortfall of pi in (62) is not binding on any of the nodes

in the default set. In other words, the shocks are large enough to generate defaults, but not so large as12A corresponding comparison is possible using partial recoveries at default, as in Rogers and Veraart (2012). This leads

to qualitatively similar results, provided claims on other banks are kept to a realistic fraction of a bank’s total assets.

24

to entirely wipe out asset value at any node. In this case, we have

sD = (1 + γ)[sDAD − wD + xD].

If we further assume that ID − (1 + γ)AD is invertible, then

sD = (1 + γ)(xD − wD)[ID − (1 + γ)AD]−1

and the systemic shortfall is

S(x) = sD · uD(γ) = (1 + γ)(xD − wD) · uD(γ), (63)

where the modified node depth vector uD(γ) is given by [ID − (1 + γ)AD]−1 · 1D. If (1 + γ)AD has

spectral radius less than 1, then

uD(γ) = [ID + (1 + γ)AD + (1 + γ)2A2D + · · · ] · 1D.

The representation in (63) reveals two effects from introducing bankruptcy costs. The first is an imme-

diate or local impact of multiplying xi−wi by 1 + γ; every element of sD is positive, so the outer factor

of 1 + γ increases the total shortfall. The second and more important effect is through the increased

node depth. In particular, letting αD denote the cohesiveness of D as before, we can now lower-bound

the depth of each node by 1/[1 − (1 + γ)αD].13 This makes explicit how bankruptcy costs deepen the

losses at defaulted nodes and increase total losses to the system. By the argument in Theorem 3, we get

the following comparison of losses with and without interconnections:

Corollary 4. In the setting of Theorem 3, if we introduce bankruptcy costs satisfying (1 + γ)β+ < 1,

thenL

Lo≤ 1 +

∑δici

[1− (1 + γ)β+]∑

ci.

In the EBA data in the appendix, we have β+ = 0.43. If we set γ at the rather large value of 0.5 and

continue to assume δi ≤ 0.01, the corollary gives an upper bound of 1.042. In other words, even with

large bankruptcy costs, the additional expected loss attributable to the network is at most 4.2%.

5.2 Example

We saw previously that in the example of Figure 2(a) we need a shock greater than 80 to the outside

assets of the central node in order to have contagion to all other nodes. Under a beta distribution13If the spectral radius of (1 + γ)A is less than 1, then (1 + γ)αD < 1.

25

with parameters p = 1 and q ≥ 1, this has probability (1 − 80/150)q = (7/15)q. If we assume i.i.d.

proportional shocks to the outside assets of all nodes, then all nodes default directly with probability

[(1− w1/c1) · · · (1− w5/c5)]q = [(14/15) · 0.94]q ≈ 0.61q > (7/15)q.

Thus contagion is weak.

Now introduce a bankruptcy cost factor of γ = 0.5 and consider a shock of 57 or larger to the

central node. The shock creates a shortfall of at least 47 at the central node that gets magnified by

50% to a shortfall of at least 70.5 after bankruptcy costs. The central node’s total liabilities are 140,

so the result is that each peripheral node receives less than half of what it is due from the central

node, and this is sufficient to push every peripheral node into default. Thus, with bankruptcy costs, the

probability of contagion is at least (1− (57/150))q = 0.62q, which is now greater than the probability of

direct defaults through independent shocks. A similar comparison holds with truncated exponentially

distributed shocks.

This example illustrates how bankruptcy costs increase the probability of contagion. However, it is

noteworthy that γ needs to be quite large to overcome the effect of weak contagion.

6 Confidence, credit quality and mark-to-market losses

In the previous section, we demonstrated how bankruptcy costs can amplify losses. In fact, a borrower’s

deteriorating credit quality can create mark-to-market losses for a lender well before the point of default.

Indeed, by some estimates, these types of losses substantially exceeded losses from outright default during

2007-2009. We introduce a mechanism for adding this feature to a network model and show that it too

magnifies contagion. See Harris, Herz, and Nissim (2012) for a broad discussion of accounting practices

as sources of systemic risk.

The mechanism we introduce is illustrated in Figure 3, which shows how the value of node i’s total

liabilities changes with the level z of node i’s assets. The figure shows the special case of a piecewise

linear relationship. More generally, let r(z) be the reduced value of liabilities at a node as a function of

asset level z, where r(z) is increasing, continuous, and

0 ≤ r(z) ≤ pi, for z < (1 + k)pi;

r(z) = pi, for z ≥ (1 + k)pi.

Let R(z1, . . . , zn) = (r(z1), . . . , r(zn)). Given a shock x = (x1, . . . , xn), the clearing vector p(z) solves

p(x) = R(c + p(x)A + w − x).

Our conditions on r ensure the existence of such a fixed point by an argument similar to the one in

Section 2.2.

26

ip

ip

[1 (1 )] ik p zη η− + +

ipk)1( +

( )( ) 1ik p zγ η γ− + + +

Assets z

Value of Liabilities ( )r z

Figure 3: Liability value as a function of asset level in the presence of bankruptcy costs and creditquality.

The effect of credit quality deterioration begins at a much higher asset level of z = (1 + k)pi than

does default. Think of k as measuring a capital cushion: node i’s credit quality is impaired once its

net worth (the difference between its assets and liabilities) falls below the cushion. At this point, the

value of i’s liabilities begins to decrease, reflecting the mark-to-market impact of i’s deteriorating credit

quality.

In the example of Figure 2(a), suppose the central node is at its minimum capital cushion before

experiencing a shock; in other words (1 + k)pi = 150, which implies that k = 1/14. A shock of 5 to the

central node’s outside assets reduces the total value of its liabilities by 5η, so each peripheral node incurs

a mark-to-market loss of 5η/14. In contrast, in the original model with η = 0, the peripheral nodes do

not experience a loss unless the shock to the central node exceeds 10.

As this example illustrates, mark-to-market losses from credit deterioration and bankruptcy costs

operate in qualitatively different ways, even though both increase overall losses. Bankruptcy costs

amplify the propagation of shortfall. In contrast, accounting for credit quality effectively increases the

linkages between nodes. It does so by propagating losses at higher levels of asset values, rather than

by amplifying the losses as they are propagated. Shocks can also be propagated through prices due to

common exposures or fire sales that go beyond the network of direct obligations, as in Allen and Gale

(2000), Cifuentes, Ferrucci, and Shin (2005), and Caccioli et al. (2012).

7 Concluding Remarks

In this paper we have argued that it is relatively difficult to generate contagion solely through spillover

losses in a network of payment obligations. For contagion to occur, a shock to one node must lead to

losses at another set of nodes sufficiently large to wipe out their initial equity (or net worth). This means

that either the initial shock must be large (and therefore improbable) relative to the net worth of the

infecting node, or that the net worth of the infected nodes must be small. More generally, the probability

27

of contagion depends on the comparative net worths of the various nodes, their level of leverage, and

the extent to which the infecting node has obligations to the rest of the financial sector. As Theorems

1 and 2 show, one can compare the probability of contagion and direct default under a wide range of

shock distributions without knowing the topological details of the network.

The network structure matters more for the amplification effect, in which losses among defaulting

nodes multiply because of their obligations to one another. The degree of amplification is captured by

the concept of node depth, which is the expected number of periods it takes to exit from the default set

from a given starting point. This clearly depends on the specific structure of the interbank obligations,

and is dual to the concept of eigenvector centrality in the networks literature. As Theorem 3 shows one

can place an upper bound on the amplification effect based on banks’ maximum connectedness with the

rest of the financial system (the parameter β+) without knowing any details of the connections within

the system.

The network structure takes on added importance for both contagion and amplification once we

introduce bankruptcy costs and mark-to-market reductions in credit quality. Bankruptcy costs steepen

the losses at defaulted nodes, thereby increasing the likelihood that defaults will spread to other nodes.

These losses are further amplified by feedback effects, thus increasing the system-wide loss in value. By

contrast, reductions in credit quality have the effect of marking down asset values in advance of default.

This process is akin to a slippery slope: once some node suffers a deterioration in its balance sheet,

its mark-to-market value decreases, which reduces the value of the nodes to which it has obligations,

causing their balance sheets to deteriorate. The result can be a system-wide reduction in value that was

triggered solely by a loss of confidence rather than an actual default.

References

[1] Acemoglu, D., Ozdaglar, A., and Tahbaz-Salehi, A. (2013). Systemic Risk and Stability in Finan-

cial Networks. Working Paper 18727, National Bureau of Economic Research, Cambridge, Mas-

sachusetts.

[2] Allen, F., and Babus, A. (2009). Networks in Finance. In The Network Challenge: Strategy, Profit,

and Risk in an Interlinked World (eds. P. Kleindorfer, Y. Wind & R. Gunther), Wharton School

Publishing, Upper Saddle River, New Jersey.

[3] Allen, F., and Carletti, E. (2011). New Theories to Underpin Financial Reform, Journal of Financial

Stability, in press.

[4] Allen, F., and Gale, G. (2000). Financial Contagion, Journal of Political Economy 108 1–33.

28

[5] Bank of England (2011). Financial Stability Report 29. Available at www.bankofengland.co.uk.

[6] Barlow, R.E., and Proschan, F. (1975). Statistical Theory of Reliability and Life Testing, Holt,

Rinehart, and Winston, New York.

[7] Battiston, S., Delli Gatti, D., Gallegati, M., Greenwald, B., and Stiglitz, J.E. (2012). Default

Cascades: When Does Risk Diversification Increase Stability? Journal of Financial Stability 8,

138-149.

[8] Caccioli, F., Shrestha, M., Moore, C., and Farmer, J.D. (2012). Stability Analysis of Financial

Contagion Due to Overlapping Portfolios. Working Paper 2012-10-018, Santa Fe Institute.

[9] Cifuentes, R., Ferrucci, G., and Shin, H.S. (2005). Liquidity Risk and Contagion, Journal of the

European Economic Association 3, 556–566.

[10] Cont, R., Moussa, A., and Santos, E. (2010). Network Structure and Systemic Risk in Banking

Systems. Available at SSRN: http://ssrn.com/abstract=1733528.

[11] Degryse, H., and Nguyen, G. (2004). Interbank Exposures: An Empirical Examination of Systemic

Risk in the Belgian Banking System. Working paper, Belgian National Bank.

[12] Demange, G. (2012). Contagion in Financial Networks: A Threat Index. Working Paper 2012-02,

Paris School of Economics, Paris.

[13] Diamond, D.W., and Dybvig, P. (1983). Bank Runs, Deposit Insurance, and Liquidity, Journal of

Political Economy 91, 401–419.

[14] Eisenberg, L., and Noe, T.H. (2001). Systemic Risk in Financial Systems, Management Science 47,

236–249.

[15] Elliott, M., Golub, B., and Jackson, M.O. (2013) Financial Networks and Contagion. Working

Paper, California Institute of Technology, Pasadena, California.

[16] Elsinger, H., Lehar, A., and Summer, M. (2006). Risk Assessment for Banking Systems, Management

Science 52, 1301–1314.

[17] European Banking Authority (2011), EU-Wide Stress Test Results. Available at www.eba.europa.eu.

[18] Furfine, C. (2003). Interbank Exposures: Quantifying the Risk of Contagion, Journal of Money,

Credit and Banking 35, 11–128.

[19] Gai, P., Haldane, A., and Kapadia, S. (2011). Complexity, Concentration, and Contagion, Journal

of Monetary Economics 58, 453–470.

29

[20] Gai, P., and Kapadia, S. (2010). Contagion in Financial Networks, Proceedings of the Royal Society

A 466, 2401-2423.

[21] Georg, C.-P. (2011). The Effect of the Interbank Network Structure on Contagion and Common

Shocks, Deutsche Bundesbank Discussion Paper Series 2: Banking and Financial Studies, No.

12/2011.

[22] Goodhart, C.A.E., Sunirand, P., and Tsomocos, D. P. (2004). A Model to Analyse Financial

Fragility: Applications, Journal of Financial Stability 1, 1-30.

[23] Haldane, A. G., and May, R. M. (2011). Systemic Risk in Banking Ecosystems, Nature 469, 351-355.

[24] Harris, T. S., Herz, R.H., and Nissim, D. (2012) Accounting’s Role in the Reporting, Creation,

and Avoidance of Systemic Risk in Financial Institutions. In The Handbook of Systemic Risk (eds.

J.P-Fouqe and J. Langsam), Cambridge University Press.

[25] International Monetary Fund (2012). Global Financial Stability Report: Restoring Confidence and

Progressing on Reforms. Available at www.imf.org/external/pubs/ft/gfsr.

[26] Liu, M., and Staum, J. (2012). Systemic Risk Components in a Network Model of Contagion.

Available at SSRN: http://ssrn.com/abstract=1726107.

[27] Marshall, A.W., and Olkin, I. (1974). Majorization in Multivariate Distributions. The Annals of

Statistics 2, 1189-1200.

[28] Morris, S. (2000). Contagion. Review of Economic Studies 67, 57-78.

[29] Newman, M. (2010). Networks: An Introduction. Oxford University Press, USA.

[30] Nier, E., Yang, J., Yorulmazer, T., and Alentorn, A. (2007) Network Models and Financial Stability,

Journal of Economic Dynamics and Control 31, 2033-2060.

[31] Office of Financial Research (2012). Annual Report. Available at www.treasury.gov/ofr.

[32] Ramos Romer, H.M., and Sordo Diaz, M.A. (2001) The Proportional Likelihood Ratio Order and

Applications. QUESTIIO 25, 211-223.

[33] Rogers, L.C.G., and Veraart, L.A.M. (2012). Failure and Rescue in an Interbank Network. Manage-

ment Science, in press.

[34] Shin, H. (2012) Global Banking Glut and Loan Risk Premium. Mundell-Fleming Lecture, Interna-

tional Monetary Fund, Washington, D.C.

30

[35] Shleifer, A., and Vishny, R. (2011). Fire Sales in Finance and Economics. Journal of Economic

Perspectives 25, 29–48.

[36] Tasche, D. (2008). The Vasicek Distribution. Available at wwwold-m4.ma.tum.de/pers/tasche/.

[37] Upper, C., (2011). Simulation Methods to Assess the Danger of Contagion in Interbank Markets,

Journal of Financial Stability 7, 111-125.

[38] Upper, C., and Worms, A. (2002). Estimating Bilateral Exposures in the German Interbank Market:

Is There a Danger of Contagion? Discussion Paper 09, Deutsche Bundesbank.

[39] Vuillemey, G., and Peltonen, T.A. (2013) Sovereign Credit Events and Their Spillovers to the

European Banking System — The Interplay Between Sovereign Bonds and CDS Holdings. Working

paper, European Central Bank.

[40] Wirch, J.L. (1999) Coherent Beta Risk Measures for Capital Requirements. Dissertation, Depart-

ment of Statistics and Actuarial Sciences, University of Waterloo, Waterloo, Ontario, Canada.

[41] Yellen, J. (2013) Interconnectedness and Systemic Risk: Lessons from the Financial Crisis and

Policy Implications. Board of Governors of the Federal Reserve System, Washington, D.C.

31

Appendix: Application to European Banks

To provide some insight into how our theoretical results apply in practice, we draw on data from banks

that participated in the European Banking Authority’s (EBA) 2011 stress test. Detailed information on

interbank exposures needed to calibrate a full network model is not publicly available. But our results

do not depend on detailed network structure, so the information disclosed with the results of the stress

test give us most of what we need.

Ninety banks in 21 countries participated in the stress test. For each, the EBA reports total assets

and equity values as of the end of 2010. In addition, the EBA reports each bank’s total exposure at

default (EAD) to other financial institutions. The EAD measures a bank’s total claims on all other

banks, so we take this as the size of each bank’s in-network assets. Subtracting this value from total

assets gives us ci, the size of the bank’s outside assets. For wi we use the equity values reported by the

EBA, which then allows us to calculate λi = ci/wi.

The only remaining parameter we need is βi, the fraction of a bank’s liabilities owed to other banks.

This information is not included in the EBA summary, nor is it consistently reported by banks in their

financial statements. As a rough indication, we assume that each bank’s in-network liabilities equal its

in-network assets (though we will see that our results are fairly robust to this assumption).14 This gives

us βi = EAD/(assets-equity).

Some of the smallest banks have problematic data, so as a simple rule we omit the ten smallest. We

also omit any countries with only a single participating bank. This leaves us with 76 banks, of which

the 50 largest are included in Table 1. For all 76 banks, Table 2 includes assets, EAD, and equity as

reported by the EBA (in millions of euros), and our derived values for ci, wi, βi, λi. The banks are listed

by asset size in Table 1 and grouped by country in Table 2.

In Table 1, we examine the potential for contagion from the failure of each of the five largest banks,

BNP Paribas, Deutsche Bank, HSBC, Barclays, and Credit Agricole. Taking each of these in turn as

the triggering bank, we then take the default set D to be consecutive pairs of banks. The first default

set under BNP Paribas consists of Deutsche Bank and HSBC, the next default set consists of HSBC and

Barclays, and so on.

Under “WR” (weak ratio), we report the ratio of the left side of inequality (20) to the right side.

Contagion is weak whenever this ratio exceeds 1, as it does in most cases in the table. Contagion fails to

be weak only when the banks in the default set are much smaller than the triggering bank. Moreover,

the ratio reported for each bank shows how much greater βi would have to be to reverse the direction

of inequality (20). For example, the first ratio listed under BNP Paribas, corresponding to the default

14Based on Federal Reserve Release H.8, the average value of βi for commercial banks in the U.S. is about 3%, so ourestimates for European banks would appear to be conservative.

32

of Deutsche Bank and HSBC, is 18.64, based on a βi value of 4.6% (see Table 2). This tells us that

the βi value would have to be at least 18.64 x 4.6% = 85.7% for the weak contagion condition to be

violated. In this sense, the overall pattern in Table 1 is robust to our estimated values of βi. Expanding

the default sets generally makes contagion weaker because of the relative magnitudes of wi and λ−1i ; see

(20) and Table 2.

Under “LR” (for likelihood ratio), we report the relative probability of failure through independent

direct shocks and through contagion, calculated as the ratio of the right side of (22) to the left side.

This is the ratio of probabilities under the assumption of a uniform distribution p = q = 1, which

is conservative. An asterisk indicates that the ratio is infinite because default through contagion is

impossible.15 The value of 6.68 reported under BNP Paribas for the default set consisting of Banco

Santander and Societe Generale indicates that the probability of default through independent shocks is

6.68 times more likely than default through contagion. Raising the LR values in the table to a power of

q > 1 gives the corresponding ratio of probabilities under a shock distribution having parameters p = 1

and q.16

Table 2 reports a similar analysis by country. Within each country, we first consider the possibility

that the failure of a bank causes the next two largest banks to default – the next two largest banks

constitute the default set D. (In the case of Belgium, there is only one bank to include in D.) For each

bank, the column labeled “Ratio for Next Two Banks” reports the ratio of the left side of inequality

(20) to the right side. As in Table 1, contagion is weak whenever this ratio exceeds 1. The table shows

that the ratio is greater than 1 in every case. As in Table 1, the magnitude of the ratio also tells us how

much larger βi would have to be to reverse the inequality in (20).

The last column of the table reports the corresponding test for weak contagion but now holding D

fixed as the two smallest banks in each country group. We now see several cases in which the ratio is

less than 1 – for example, when we take Deutsche Bank to be the triggering bank and the two smallest

banks in the German group (Landesbank Berlin and DekaBank Deutsch Girozentrale) as the default set

we get a ratio of 0.5 in the last column.

The results in Tables 1 and 2 reflect the observations we made following Theorem 1 about the relative

magnitudes needed for the various model parameters in order that contagion not be weak, and they show

that the parameter ranges implying weak contagion are indeed meaningful in practice.

15In deriving (32), we use the bound in (38), which is conservative. Thus, LR must be greater than 1 whenever WR is,and WR can be finite even when LR is infinite and default through contagion is impossible.

16We noted previously that a beta distribution can be used to approximate the Gaussian copula loss distribution usedin Basel capital standards. We find that the best fit occurs at values of q around 20 or larger. Raising the likelihood ratiosin the table to the qth power thus has a major impact.

33

WR LR WR LR WR LR WR LR WR LR

BNP PARIBAS

DEUTSCHE BANK AG

HSBC HOLDINGS plc 18.64 *

BARCLAYS plc 18.63 * 9.21 3.89

CREDIT AGRICOLE 17.83 * 8.81 2.01 8.27 1.88 27.23

BANCO SANTANDER S.A. 14.91 * 7.37 1.98 6.92 1.85 24.62 *

SOCIETE GENERALE 12.52 6.68 6.19 1.62 5.81 1.55 20.67 * 13.66 15.32

LLOYDS BANKING GROUP plc 11.23 7.70 5.55 1.63 5.21 1.56 18.55 * 12.26 23.52

BPCE 11.29 12.77 5.58 1.70 5.24 1.62 18.65 * 12.32 *

ING BANK NV 10.32 3.44 5.10 1.46 4.79 1.41 17.05 * 11.27 4.57

UNICREDIT S.p.A 9.55 4.06 4.72 1.49 4.43 1.44 15.77 * 10.42 5.90

COMMERZBANK AG 8.47 3.33 4.18 1.43 3.93 1.38 13.98 * 9.24 4.39

RABOBANK NEDERLAND 6.96 2.48 3.44 1.33 3.23 1.29 11.49 * 7.59 2.94

ROYAL BANK OF SCOTLAND GROUP plc 6.06 * 2.99 1.67 2.81 1.58 10.01 * 6.61 *

INTESA SANPAOLO S.p.A 5.70 72.67 2.82 1.63 2.64 1.54 9.41 * 6.22 *

DEXIA 4.59 1.79 2.27 1.19 2.13 1.16 7.58 5.19 5.01 1.97

Nordea Bank AB (publ) 4.52 1.57 2.23 1.15 2.10 1.13 7.46 2.96 4.93 1.68

BANCO BILBAO VIZCAYA ARGENTARIA  5.24 1.85 2.59 1.21 2.43 1.19 8.65 5.82 5.72 2.05

DANSKE BANK 4.47 1.66 2.21 1.16 2.08 1.14 7.39 3.72 4.88 1.80

ABN AMRO BANK NV 3.91 1.33 1.93 1.09 1.81 1.08 6.46 1.85 4.27 1.38

Landesbank Baden‐Württemberg 3.36 1.23 1.66 1.06 1.56 1.05 5.55 1.57 3.67 1.27

Hypo Real Estate Holding AG, Münche 3.00 1.15 1.48 1.03 1.39 1.03 4.96 1.34 3.28 1.17

BFA‐BANKIA 3.38 1.21 1.67 1.05 1.57 1.04 5.59 1.49 3.69 1.24

DZ BANK AG Dt. Zentral‐Genossenscha 2.83 1.21 1.40 1.04 1.31 1.03 4.67 1.54 3.08 1.25

Bayerische Landesbank 2.39 1.16 1.18 1.02 1.11 1.01 3.94 1.42 2.61 1.20

KBC BANK 2.73 1.24 1.35 1.04 1.27 1.03 4.50 1.63 2.98 1.28

CAJA DE AHORROS Y PENSIONES DE BA 3.03 1.24 1.50 1.05 1.41 1.04 5.01 1.63 3.31 1.29

BANCA MONTE DEI PASCHI DI SIENA S 2.96 1.17 1.46 1.04 1.38 1.03 4.89 1.40 3.23 1.20

Svenska Handelsbanken AB (publ) 2.61 1.13 1.29 1.02 1.21 1.02 4.31 1.30 2.85 1.15

Norddeutsche Landesbank ‐GZ 2.13 1.09 1.05 1.00 0.99 1.00 3.52 1.22 2.33 1.11

Skandinaviska Enskilda Banken AB (pub 1.96 1.09 0.97 1.00 0.91 0.99 3.24 1.25 2.14 1.11

DnB NOR Bank ASA 2.26 1.16 1.12 1.02 1.05 1.01 3.73 1.44 2.46 1.20

Erste Bank Group (EBG) 2.22 1.17 1.10 1.01 1.03 1.01 3.66 1.47 2.42 1.21

WestLB AG, Düsseldorf 1.91 1.10 0.94 1.00 0.88 0.99 3.15 1.27 2.08 1.12

Swedbank AB (publ) 1.82 1.07 0.90 0.99 0.84 0.99 3.00 1.19 1.98 1.09

Nykredit 1.98 1.10 0.98 1.00 0.92 0.99 3.27 1.26 2.16 1.12

BANK OF IRELAND 1.76 1.08 0.87 0.99 0.82 0.98 2.91 1.24 1.92 1.10

HSH Nordbank AG, Hamburg 1.63 1.06 0.80 0.98 0.76 0.98 2.69 1.18 1.78 1.08

BANCO POPOLARE ‐ S.C. 1.58 1.05 0.78 0.98 0.73 0.98 2.61 1.15 1.73 1.06

Landesbank Berlin AG 1.21 1.03 0.60 0.96 0.56 0.95 1.99 1.13 1.32 1.04

ALLIED IRISH BANKS PLC 1.12 1.01 0.55 0.96 0.52 0.96 1.85 1.10 1.22 1.03

Raiffeisen Bank International (RBI) 1.23 1.03 0.61 0.96 0.57 0.95 2.04 1.14 1.35 1.05

UNIONE DI BANCHE ITALIANE SCPA (U 1.22 1.04 0.60 0.95 0.57 0.94 2.02 1.19 1.33 1.06

DekaBank Deutsche Girozentrale, Fran 1.19 1.02 0.59 0.96 0.55 0.96 1.96 1.12 1.30 1.04

BANCO POPULAR ESPAÑOL, S.A. 1.21 1.03 0.60 0.96 0.56 0.96 2.00 1.12 1.32 1.04

CAIXA GERAL DE DEPÓSITOS, SA 1.27 1.04 0.63 0.96 0.59 0.95 2.10 1.18 1.39 1.06

NATIONAL BANK OF GREECE 1.25 1.05 0.62 0.95 0.58 0.94 2.07 1.21 1.37 1.07

BANCO COMERCIAL PORTUGUÊS, SA (B 1.21 1.03 0.60 0.95 0.56 0.95 2.00 1.15 1.32 1.05

BANCO DE SABADELL, S.A. 1.08 1.01 0.53 0.96 0.50 0.96 1.78 1.07 1.17 1.02

EFG EUROBANK ERGASIAS S.A. 1.01 1.00 0.50 0.95 0.47 0.95 1.66 1.07 1.10 1.01

ESPÍRITO SANTO FINANCIAL GROUP, SA 0.92 0.99 0.46 0.94 0.43 0.94 1.52 1.07 1.01 1.00

From BNP Paribas From Deutsche Bank From HSBC From Barclays From Credit Agricole

Table 1: Results based on 2011 EBA stress test data. We consider contagion from each of the five largest banksacross the top to pairs of banks in consecutive rows. A weak ratio (WR) value greater than 1 indicates thatcontagion is weak. Each LR value is a likelihood ratio for default through independent shocks to default throughcontagion, assuming a uniform fractional shock.

34

c_i w_i beta_i lambda_i

Assets EAD Outside Assets Equity

EAD/(Assets‐

Equity)

Outside 

Assets/Equity

Ratio for Next 

Two Banks

Ratio for Last 

Two Banks

AT001 Erste Bank Group (EBG) 205,938             25,044           180,894            10,507       12.8% 17.2

AT002 Raiffeisen Bank International (RBI) 131,173             30,361           100,812            7,641         24.6% 13.2

AT003 Oesterreichische Volksbank AG 44,745               10,788           33,957              1,765         25.1% 19.2 3.4

BE004 DEXIA 548,135             228,211         319,924            17,002       43.0% 18.8

BE005 KBC BANK 276,723             23,871           252,852            11,705       9.0% 21.6 35.7

CY006 MARFIN POPULAR BANK PUBLIC CO LTD 42,580               7,907             34,673              2,015         19.5% 17.2

CY007 BANK OF CYPRUS PUBLIC CO LTD 41,996               7,294             34,702              2,134         18.3% 16.3 1.6

DE017 DEUTSCHE BANK AG 1,905,630         194,399         1,711,231         30,361       10.4% 56.4 0.5

DE018 COMMERZBANK AG 771,201             138,190         633,011            26,728       18.6% 23.7 0.8

DE019 Landesbank Baden‐Württemberg 374,413             133,906         240,507            9,838         36.7% 24.4 2.5 1.0

DE020 DZ BANK AG Dt. Zentral‐Genossenschaf 323,578             135,860         187,718            7,299         43.0% 25.7 1.9 1.1

DE021 Bayerische Landesbank 316,354             97,336           219,018            11,501       31.9% 19.0 2.4 1.3

DE022 Norddeutsche Landesbank ‐GZ 228,586             91,217           137,369            3,974         40.6% 34.6 2.5 1.6

DE023 Hypo Real Estate Holding AG, München 328,119             29,084           299,035            5,539         9.0% 54.0 3.0 3.3

DE024 WestLB AG, Düsseldorf 191,523             58,128           133,395            4,218         31.0% 31.6 3.6 2.2

DE025 HSH Nordbank AG, Hamburg 150,930             9,532             141,398            4,434         6.5% 31.9 5.2 9.7

DE027 Landesbank Berlin AG 133,861             49,253           84,608              5,162         38.3% 16.4 2.6

DE028 DekaBank Deutsche Girozentrale, Frank 130,304             41,255           89,049              3,359         32.5% 26.5 9.7

DK008 DANSKE BANK 402,555             75,894           326,661            14,576       19.6% 22.4 0.4

DK011 Nykredit 175,888             8,597             167,291            6,633         5.1% 25.2 2.7

DK009 Jyske Bank 32,752               4,674             28,078              1,699         15.1% 16.5 1.4

DK010 Sydbank 20,238               3,670             16,568              1,231         19.3% 13.5 2.7

ES059 BANCO SANTANDER S.A. 1,223,267         51,407           1,171,860         41,998       4.4% 27.9 0.4

ES060 BANCO BILBAO VIZCAYA ARGENTARIA S 540,936             110,474         430,462            24,939       21.4% 17.3 0.2

ES061 BFA‐BANKIA 327,930             39,517           288,414            13,864       12.6% 20.8 7.4 0.6

ES062 CAJA DE AHORROS Y PENSIONES DE BA 275,856             5,510             270,346            11,109       2.1% 24.3 3.2 3.6

ES064 BANCO POPULAR ESPAÑOL, S.A. 129,183             14,810           114,373            6,699         12.1% 17.1 5.2 1.5

ES065 BANCO DE SABADELL, S.A. 96,703               3,678             93,025              3,507         3.9% 26.5 19.6 5.5

ES066 CAIXA D'ESTALVIS DE CATALUNYA, TAR 76,014               8,219             67,795              3,104         11.3% 21.8 6.1 2.7

ES067 CAIXA DE AFORROS DE GALICIA, VIGO,  73,319               2,948             70,371              2,849         4.2% 24.7 19.5 6.9

ES083 CAJA DE AHORROS DEL MEDITERRANEO 72,034               4,981             67,053              1,843         7.1% 36.4 9.5 4.2

ES071 GRUPO BANCA CIVICA 71,055               7,419             63,636              3,688         11.0% 17.3 22.9 3.0

ES068 GRUPO BMN 69,760               7,660             62,101              3,304         11.5% 18.8 13.6 2.9

ES063 EFFIBANK 54,523               4,124             50,399              2,656         8.0% 19.0 8.5 5.2

ES069 BANKINTER, S.A. 53,476               2,141             51,335              1,920         4.2% 26.7 7.5 9.5

ES070 CAJA ESPAÑA DE INVERSIONES, SALAM 45,656               7,235             38,422              2,076         16.6% 18.5 11.5 3.2

ES075 GRUPO BBK 44,628               1,924             42,704              2,982         4.6% 14.3 19.9 10.7

ES072 CAJA DE AHORROS Y M.P. DE ZARAGOZ 42,716               1,779             40,938              2,299         4.4% 17.8 6.9 11.5

ES073 MONTE DE PIEDAD Y CAJA DE AHORRO 34,263               2,599             31,664              2,501         8.2% 12.7 19.4 8.2

ES074 BANCO PASTOR, S.A. 31,135               1,665             29,470              1,395         5.6% 21.1 18.1 12.5

ES076 CAIXA D'ESTALVIS UNIO DE CAIXES DE M 28,310               1,802             26,508              1,065         6.6% 24.9 11.8 11.6

ES077 CAJA DE AHORROS Y M.P. DE GIPUZKOA 20,786               285                20,501              1,935         1.5% 10.6 14.2

ES078 GRUPO CAJA3 20,144               1,926             18,218              1,164         10.1% 15.7 11.6

FR013 BNP PARIBAS 1,998,157         90,328           1,907,829         55,352       4.6% 34.5 11.1

FR014 CREDIT AGRICOLE 1,503,621         83,713           1,419,908         46,277       5.7% 30.7 12.2

FR016 SOCIETE GENERALE 1,051,323         100,013         951,310            27,824       9.8% 34.2 13.9

FR015 BPCE 1,000,695         34,983           965,712            31,943       3.6% 30.2 12.2

GB089 HSBC HOLDINGS plc 1,783,199         212,092         1,571,107         86,900       12.5% 18.1 3.5

GB090 BARCLAYS plc 1,725,709         53,873           1,671,836         46,232       3.2% 36.2 12.3

GB091 LLOYDS BANKING GROUP plc 1,006,082         29,233           976,849            47,984       3.1% 20.4 6.6

GB088 ROYAL BANK OF SCOTLAND GROUP plc 607,351             105,506         501,845            58,982       19.2% 8.5 12.3

GR031 NATIONAL BANK OF GREECE 118,832             8,608             110,224            8,153         7.8% 13.5 6.0

GR030 EFG EUROBANK ERGASIAS S.A. 85,885               3,838             82,047              4,296         4.7% 19.1 12.9

GR032 ALPHA BANK 66,798               3,492             63,306              5,275         5.7% 12.0 8.9 14.4

GR033 PIRAEUS BANK GROUP 57,680               1,581             56,099              3,039         2.9% 18.5 16.5

GR034 AGRICULTURAL BANK OF GREECE S.A. ( 31,221               1,657             29,564              792             5.4% 37.3 14.4

IE038 BANK OF IRELAND 156,712             17,254           139,458            7,037         11.5% 19.8 4.9

IE037 ALLIED IRISH BANKS PLC 131,311             11,277           120,034            3,669         8.8% 32.7

IE039 IRISH LIFE AND PERMANENT 46,743               6,127             40,616              1,681         13.6% 24.2 4.9

IT041 UNICREDIT S.p.A 929,488             106,707         822,781            35,702       11.9% 23.0 1.3

IT040 INTESA SANPAOLO S.p.A 576,962             109,909         467,053            26,159       20.0% 17.9 1.4

IT042 BANCA MONTE DEI PASCHI DI SIENA S.p 244,279             12,074           232,205            6,301         5.1% 36.9 4.2 10.4

IT043 BANCO POPOLARE ‐ S.C. 140,043             7,602             132,440            5,474         5.6% 24.2 2.0

IT044 UNIONE DI BANCHE ITALIANE SCPA (UB 130,559             19,793           110,766            6,559         16.0% 16.9 10.4

NL047 ING BANK NV 933,073             111,756         821,317            30,895       12.0% 26.6 2.5

NL048 RABOBANK NEDERLAND 607,483             37,538           569,945            27,725       6.2% 20.6 7.2

NL049 ABN AMRO BANK NV 379,599             29,196           350,403            11,574       7.7% 30.3 5.1

NL050 SNS BANK NV 78,918               388                78,530              1,782         0.5% 44.1 7.2

PT053 CAIXA GERAL DE DEPÓSITOS, SA 119,318             14,221           105,097            6,510         11.9% 16.1 5.0

PT054 BANCO COMERCIAL PORTUGUÊS, SA (B 100,010             7,690             92,320              3,521         7.7% 26.2 8.5

PT055 ESPÍRITO SANTO FINANCIAL GROUP, SA 85,644               8,690             76,955              4,520         10.1% 17.0 7.1

PT056 Banco BPI, SA 43,826               5,463             38,364              2,133         12.5% 18.0 8.5

SE084 Nordea Bank AB (publ) 542,853             61,448           481,405            19,103       11.3% 25.2 3.7

SE085 Skandinaviska Enskilda Banken AB (pub 212,240             25,955           186,284            9,604         12.2% 19.4 9.0

SE086 Svenska Handelsbanken AB (publ) 240,202             20,870           219,332            8,209         8.7% 26.7 3.8

SE087 Swedbank AB (publ) 191,365             17,358           174,007            7,352         9.1% 23.7 9.0

Bank Number and Name

Table 2: Results based on 2011 EBA stress test data. In the last two columns, a ratio greater than 1 indicatesweak contagion from a bank to the next two largest banks in the same country and to the two smallest banks inthe country group, respectively. 35